• Aucun résultat trouvé

Numerical Simulations of Kinetic Models for chemotaxis

N/A
N/A
Protected

Academic year: 2021

Partager "Numerical Simulations of Kinetic Models for chemotaxis"

Copied!
18
0
0

Texte intégral

(1)

HAL Id: hal-00798822

https://hal.archives-ouvertes.fr/hal-00798822

Submitted on 10 Mar 2013

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Numerical Simulations of Kinetic Models for chemotaxis

Francis Filbet, Chang Yang

To cite this version:

(2)

NUMERICAL SIMULATIONS OF KINETIC MODELS FOR CHEMOTAXIS

FRANCIS FILBET AND CHANG YANG

Abstract. We present a new algorithm based on a Cartesian mesh for the numerical ap-proximation of kinetic models for chemosensitive movements set in an arbitrary geometry. We investigate the influence of the geometry on the collective behavior of bacteria described by a kinetic equation interacting with nutrients and chemoattractants. Numerical simula-tions are performed to verify accuracy and stability of the scheme and its ability to exhibit aggregation of cells and wave propagations. Finally some comparisons with experiments show the robustness and accuracy of such kinetic models.

Keywords. Bacterial chemotaxis, chemical signaling, kinetic theory

Contents

1. Introduction 1

2. Kinetic models for bacterial chemotaxis 2

3. Numerical resolution 4

4. Numerical examples 9

5. Conclusion 15

References 16

1. Introduction

This paper is devoted to the numerical simulation of the process of cellular spatial orga-nization driven by chemotaxis. The effective mechanism by which individual cells undergo directed motion varies among organisms. Here we are particularly interested in bacterial migration, characterized by the smallness of the cells, and their ability to move to several orders of magnitude in the attractant concentration. Several models, depending on the level of description, have been developed mathematically for the collective motion of cells. We refer to [5,24,25] for a review on parabolic, hyperbolic and kinetic models and to [3,27,28] for traveling waves drivn by growth and chemotaxis. Among them the kinetic model in-troduced by Othmer, Dunbar and Alt [2, 22], describes a population of bacteria in motion (for instance the E. Coli) in interactions with a chemoattractant concentration [10]. These cells are so small that they are not able to measure any head-to-tail gradient of the chemical concentration, and to choose directly some preferred direction of motion towards high con-centrated regions. Therefore they develop an indirect strategy to select favorable regions, by detecting a sort of time derivative in the concentration along their pathways, and to react accordingly [16].

More precisely, they undergo a jump process where free runs are followed by a reorientation phenomenon called tumble [32]. For instance it is known that E. Coli increases the time spent in running along a favourable direction [16]. This jump process can be described by two different informations. First cells switch the rotation of their flagella, from counter-clockwise,

The authors are partially supported by the European Research Council ERC Starting Grant 2009, project 239983-NuSiKiMo.

(3)

called free runs, to clockwise called reorientation or tumbling phase, and conversely. This decision is the result of a complex chain of reactions inside the cells, driven by the external concentration of the chemoattractant [29, 32]. Then bacteria select a new direction, but they are unable to choose directly a favourable direction, so they randomly choose a new orientation. During the ”run” phases a bacterium moves with a constant speed in a given direction while during a ”tumbling” event it changes direction randomly.

2. Kinetic models for bacterial chemotaxis

In the simple situation, C. S. Patlak [23] and E. F. Keller & L. A. Segel [20] considered a density of cells which interacts with two chemical substances. The cells consume nutrients which drive the migration and excrete a chemoattractant that prevents the dispersion of the population. However, this approach is not always sufficiently precise to describe the evolution of bacteria movements. Hence, this phenomenon of run and tumble can be modeled by a stochastic process called the velocity-jump process, and has been introduced by Alt [2] and further developed in [22]. A kinetic transport model to describe this velocity jump process consists to study the evolution of the bacterial population by the local density of cells f (t, x, v) ≥ 0 located in position x, at time t and swimming in the direction v ∈ V .

Here the set V of all possible velocities is bounded and symmetric in general. In two dimensions, the modulus of the speed is a constant v0, hence V = S(0, v0) circle centred in 0 with a radius v0>0.

A kinetic transport model to describe this velocity-jump process has been introduced by W. Alt [2] inspired by the Boltzmann equation [22] where the tumbles appear as scattering events and all the fluxes are explicitly introduced [27]. We consider the Boltzmann type equation

(2.1) ∂f

∂t +v ⋅ ∇xf = Q(f ) + r f, x ∈ Ω, v ∈ V, where Q(f ) is the Boltzmann type tumbling operator

Q(f ) =∫ V T (v, v ′ )f (t, x, v′)dv′ − ∫ V T (v ′ , v) f (t, x, v) dv′ .

The contribution of the tumbles is introduced with a transition (scattering) kernel T (v, v′ ) ≥0 which stands for the change of velocity from v′ to v; r is the division rate of the bacteria (r = ln 2/τ2 where τ2 is the mean doubling time). This equation is indeed a variant of the Boltzmann equation for gases, where collisions are delocalized via the secretion or consump-tion of chemical cues.

In (2.1), the transition kernel T also depends on the local concentration of chemoattractant S(t, x) and nutrient N (t, x). To estimate the respective contributions on pulse speed of the biais of the run lengths and of preferential reorientation, it is possible to split this transition kernel T (v, v′

) in two contributions, one being the tumbling rate λ(v′), and the other one the reorientation effect during tumbles K(v, v′

):

(2.2) T (v, v′

) = λ(v′)K(v, v′), with the condition

(2.3) ∫

V K(v, v ′

)dv = 1,

where the function K accounts for the persistence of the trajectories. For simplicity we consider the case of the absence of such an angular persistence, hence the turning kernel T (v, v′

)is only proportional to the tumbling rate λ(v′), i.e. K is constant. For the tumbling rate λ(v′

(4)

tumble frequency only depends on the local gradients of nutrient and attractant, both gradi-ents having independent and additive contributions. This gives

λ(v′ ) = 1 2 (λN(v ′ ) +λS(v′)) = 1 2 (ψN( D log N Dt ∣v′ ) + ψS(D log S Dt ∣v′ ) ) = 1 2(ψN(∂tlog(N ) + v ′

⋅ ∇xlog(N )) + ψS(∂tlog(S) + v′⋅ ∇xlog(S)) ) . The nutrient and chemoattractant response functions ψN and ψS are both positive and decreasing, expressing that cells are less likely to tumble (thus perform longer runs) when the external chemical signal increases. These functions are smooth and characterized by their characteristic time δ−1N and δ−1S and their tumble frequency χN and χS.

Here, we have chosen the following analytical form that encompassed these characteristics: (2.4) ψα(X) = ψ0(1 − χαtanh (X

δα

)), α = {N, S},

where ψ0is the mean tumbling frequency, and the parameters χN and χS are the modulation of tumble frequency.

In order to define completely the mathematical problem (2.1), suitable boundary conditions on ∂Ω should be applied. Here we consider wall type boundary conditions, for which emerging particles have been reflected elastically at the wall. More precisely, for x ∈ ∂Ω the smooth boundary ∂Ω is assumed to have a unit inward normal n(x) and for v ⋅ n(x) ≥ 0, we assume that at the solid boundary we have

(2.5) f (t, x, v) = R[f (t, x, v)], x ∈ ∂Ω, v ⋅ n(x) ≥ 0, with

(2.6) R[f (t, x, v)] = f (t, x, v − 2(v ⋅ n(x))n(x)).

This boundary condition (2.5) guarantees the global conservation of mass [7].

The equations describing the behaviors of nutrients density N and chemottractant S are the same as in [18]: (2.7) ⎧ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎨ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎩ ∂S ∂t − DS∆S = − a S + b∫V f (t, x, v) dv, x ∈ Ω, ∂N ∂t − DN∆N = − c N ∫V f (t, x, v) dv, x ∈ Ω,

where a, b and c are respectively the degradation rate of the chemoattractant, its production rate and the consumption rate of the nutrient by the bacteria, whereas DS and DN are the molecular diffusion coefficients. Finally, these equations are completed with homogeneous Neumann boundary conditions, i.e.

(2.8) ∇xα ⋅ n(x) = 0, α = {N, S}, x ∈ ∂Ω.

(5)

We now briefly outline the contents of the paper. In the next section, we introduce the numerical approximation of (2.1) and (2.7) and describe the numerical approximation of the boundary (2.5), (2.6), (2.8). Two points are worth mentioning here. First, we restrict ourselves to the case of specular reflection which seems to be the most appropriate for the study of bacteria. One difficulty in the approximation of kinetic models for chemotaxis, is related to the fact that it can exhibits very different phenomena as finite time blow-up, cell aggregation, wave propagations. At the discrete level, our approximation should also be able to describe a similar property. Secondly, an important step is to discretize appropriately the effect of boundary.

The final section is devoted to numerical simulations performed with the numerical scheme presented in Section 3. We investigate numerically cells aggregation, convergence to equilib-rium, and wave propagation in a bounded domain.

3. Numerical resolution

3.1. Numerical resolution of the kinetic model (2.1). We consider a computational domain [xmin, xmax] × [ymin, ymax] ×V , such that Ω ⊂ [xmin, xmax] × [ymin, ymax].

The computational domain is covered by an uniform Cartesian mesh Xh×Vh, where Xh, Vh are defined by (3.1) ⎧ ⎪ ⎪ ⎪ ⎨ ⎪ ⎪ ⎪ ⎩ Xh ∶= {x0= (xmin, ymin), . . . , xi= (xix, yiy), . . . , x(n x,ny)= (xmax, ymax) }, Vh ∶= {vj = v0(cos θj, sin θj), θj = (j + 1/2) ∆v, 0 ≤ j ≤ nv−1} .

The mesh steps are respectively ∆x = (xmax−xmin)/nx, ∆y = (ymax−ymin)/ny and ∆v = 2π/nv. Let us denote fi,jn an approximation of the distribution function f (tn, xi, vj). We introduce the following finite difference scheme

(3.2) f

n+1 i,j −fi,jn

∆t +v ⋅ ∇x,hf n

i,j = Qh(fi,jn) +r ⋅ fi,jn,

where h = (∆t, ∆x, ∆y), v ⋅ ∇x,hfi,jn is a second-order approximation [31] of the transport operator v ⋅ ∇xf , and Qh(fi,jn) is an approximation of the Boltzmann type tumbling operator Q(f ). We will now focus on searching the approximation Qh(fi,jn).

By using the trapezoidal rule, we have Qh(fi,jn) =∆v nv−1 ∑ `=0 ((Th)n+1/2i,j,` fi,`n) −∆v nv−1 ∑ `=0 ((Th)n+1/2i,`,j fi,jn), where (Th)n+1/2

i,j,` is an approximation of the transition kernel T (v, v ′

). We assume that the nutrient density N and the chemottractant S are known at time tn+1/2. Moreover with the hypothesis that K is constant, the condition (2.3) implies that K = 1/2π. Thus it remains to search an approximation of the tumbling rate, i.e. λn+1/2i,` . It is also equivalent to search the local gradients of nutrient (λN)n+1/2

i,` and attractant (λS) n+1/2 i,` . We study only the local gradients of nutrient (λN)n+1/2

i,` , since the local gradient of at-tractant (λS)n+1/2

i,` has the same expression as (λN) n+1/2

(6)

where DhNin+1/2 is a discrete time derivative and will be given in the section 3.2. Moreover we use centred difference approximation for v′

⋅ ∇xN , which yields v`⋅ ∇x,hNin+1/2=v0cos θ` Nin+1/2 x+1,iy−N n+1/2 ix−1,iy 2∆x +v0sin θ` Nin+1/2 x,iy+1−N n+1/2 ix,iy−1 2∆y .

In summary, the discrete tumbling rate reads λn+1/2i,` = 1 2((λN) n+1/2 i,` + (λS) n+1/2 i,` ). Finally we reduce the tumbling operator as follows

Qh(fi,jn) = ∆v 2π

nv−1

∑ `=0

(λn+1/2i,` fi,`n) −λn+1/2i,j fi,jn.

3.2. Calculate the chemoattractant S or the nutrient N . The equations (2.7) for nutrient density N and chemottractant S are parabolic equations with source terms depending on the distribution function of density f . We study again the discretization for nutrient density, since the one for chemottractant is similar. The Euler implicit scheme is used for time integration. Hence the scheme for nutrient density N reads

(3.3) DhNin+1/2−DNhNin+1/2= −cNin+1/2ρni, where DhNin+1/2 is an approximation of time derivative as follows

(3.4) DhNin+1/2=

Nin+1/2−Nin−1/2

∆t .

Then we use a five points finite difference scheme to discretize ∆N

(3.5) ∆hN n+1/2 i = Nin+1/2x+1,iy−2Nin+1/2 x,iy +N n+1/2 ix−1,iy ∆x2 + Nin+1/2x,iy+1−2Nin+1/2 x,iy +N n+1/2 ix,iy−1 ∆y2 .

Finally, a trapezoidal rule is applied for the integration, i.e. ρni =∆v

nv−1

∑ j=0

fi,jn.

3.3. Treatment of the boundary conditions. As we mentioned at the end of section 2, an appropriate discretization of boundary condition is important to exhibit very different phenomena. Therefore, we present respectively the numerical approximations for specular reflection boundary condition (2.6) and Neumann boundary condition (2.8).

3.3.1. Numerical approximation for specular reflection boundary condition. The specular re-flection boundary condition in 2D reads as

(3.6) R[f ](t, x, v) = f (t, x, v′),

with

v ⋅ n = −v′

⋅n ⇔ v′=v − 2(v ⋅ n)n

where x ∈ Γx=∂Ω is the point at the boundary, n is the interior normal at point x. We note that this specular reflection is just like a mirror. For example, if we follows the characteristic of the flux f (v), its reflected flow is corresponding to the velocity v′.

(7)

Thus instead of computing f at the ghost point xg, we compute f at mirror point with respect to the boundary xm=2xp−xg as follows

(3.7) f (xg, v) = f (xm, v′). x y n(xp) n ◾ ◾ ◾ ● ◯ ◯ ◯ ◾ ◾ ● ● ● ◯ ◯ ◯ xg ● ◯● ◯● ◯● ● xp P∗ 0◆ ● ● ● ● ● P∗ 1 ◆ ● ● ● ◆● ● P∗ 2 ⊡ ix− 2 ix− 1 ix ix+ 1 ix+ 2 iy− 2 iy− 1 iy iy+ 1 iy+ 2 ⊡ ⊙ ⊙xm

Figure 1. Spatially two-dimensional Cartesian mesh. ● is interior point, ◾ is ghost point,  is the point at the boundary, ◯is the point for extrapolation, the dashed line is the boundary.

The last step is to approximate f (xm, v′) using f of interior domain. Let us assume that the values of the distribution function f on the grid points in Ω are given. We first construct a stencil E composed of grid points of Ω for interpolation or extrapolation. For instance as it is shown in Figure 1, the inward normal n(xp)intersects the grid lines y = yiy, yiy+1, yiy+2 at points P∗

0, P1∗, P2∗. Then we choose the three nearest points of the cross point Pl∗, l = 0, 1, 2, in each line, i.e. marked by a large circle. From these nine points, we can build a Lagrange polynomial q2(x) ∈ Q2(R2). Therefore we evaluate the polynomial q2(x) at xm, and obtain an approximation of f at the mirror point. We distinguish two cases of mirror points. In the case that mirror point xm is surrounded by interior points, we interpolate f at mirror point xm; otherwise a WENO type extrapolation can be used to prevent spurious oscillations, which will be detailed below.

Note that in some cases, we can not find a stencil of nine interior points. For instance when the interior domain has small acute angle sharp, the normal n can not have three cross points P∗

l , l = 0, 1, 2 in interior domain, or we can not have three nearest points of the cross point P∗

(8)

which is designed to help us choose automatically between the high order accuracy and the low order but more robust extrapolation. Moreover a slightly modified version of the method was given in [14], such that the smoothness indicators are invariant with respect to the scaling of f . We now describe this method in 2D case.

The substencils Sr, r = 0, 1, 2 for extrapolation are chosen around the inward normal n such that we can construct Lagrange polynomial of degree r. For instance in Figure 1, the three substencils are respectively

⎧ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎨ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎪ ⎩ S0 = {(xix, yiy)}, S1 = {(xix−1, yiy), (xix, yiy), (xix, yiy+1), (xix+1, yiy+1)}, S2 = {(xix−1, yiy), (xix, yiy), (xix+1, yiy), (xix−1, yiy+1), (xix, yiy+1), (xix+1, yiy+1), (xix, yiy+2), (xix+1, yiy+2), (xix+2, yiy+2)}.

Once the substencils Sr are chosen, we could easily construct the Lagrange polynomials in Qr(R2) qr(x) = r ∑ m=0 r ∑ l=0 al,mxlym satisfying qr(x) = f (x), x ∈ Sr. Then the WENO extrapolation has the form

(3.8) f (x) =

2 ∑ r=0

wrqr(x), x ∈ Sr, where wr are the nonlinear weights, which are chosen to be

wr= αr ∑2s=0αs , with αr= dr (ε + βr)2 , where ε = 10−6, d0 =∆x2+∆y2, d1 = √ ∆x2+∆y2, d

2 =1 − d0−d1. The coefficients βr are the smoothness indicators determined by

β0 = ∆x2+∆y2,

and for r ≥ 1, either we take βr=0 when f (x) = 0, ∀x ∈ Sr, or we choose

βr = 1 ∑x∈Srf (x)2 ∑ 1≤∣σ∣≤r ∫ K ∣K∣∣σ∣−1(Dσqr(x))2dx, r = 1, 2,

where σ is a multi-index and K = [xp−∆x/2, xp+∆x/2] × [yp−∆y/2, yp+∆y/2] and the point xp is given by (xp, yp).

3.3.2. Numerical approximation for Neumann boundary condition. In section 3.2, we have seen that discrete Laplace operator (3.5) is the only non-locally in space term in (2.7). In Figure 2, we illustrate an example of a discrete Laplace operator near the boundary. We note that Nix,iy−1 is not known, since xg = (xix, yiy−1) is out of domain. To approximate N at

ghost point xg, we have to use the boundary condition (2.8).

(9)

Using a centered difference formula, it yields nx N (xm) −N (xg) xm−xg +ny N (xm) −N (xg) ym−yg =0,

where xg∶= (xg, yg), xm∶= (xm, ym)is the mirror point of the ghost point xg with respect to the boundary. The previous equation is equivalent to

(3.9) N (xm) =N (xg).

Therefore instead of computing N at the ghost point xg, we extrapolate N at the mirror point xm∶= (xm, ym). As shown in Figure 2, a nine points stencil S2 is found, thus we have (3.10) Nix,iy−1=N (xm) =

8 ∑ i=0

wiN (xi), xi∈ S2,

where wi is weight of extrapolation calculated, for instance, by Lagrange polynomial. Finally by injecting (3.10) into (3.5), we complete the discretization.

Remark 3.1. It is important to emphasize that the numerical scheme for the internal domain is disconnected to the numerical procedure for the treatment of the boundary. Therefore the present method can be applied to any other numerical scheme. The main point is to preserve the order of accuracy and then to eventually increase the stencil outside the domain.

x y n ◾ ◾ ◾ ◯ ◯ ◯ ◾ ◾ ◾ ◾ ● ◯ ◯ ◯ xg ● ◯● ◯● ◯● ● xp ● ● ● ● ● ● ● ● ● ● ⊡ ix− 2 ix− 1 ix ix+ 1 ix+ 2 iy− 2 iy− 1 iy iy+ 1 iy+ 2 ⊙xm

(10)

Table 1. Summary of the values used in the simulation

Parameter Value

run speed v0=25µm ⋅ s−1

mean tumbling frequency ψ0=3s−1

modulation of tumbling frequency of nutrient χN =60% modulation of tumbling frequency of chemoattractant χS=20%

stiffness of the response functions 1/δN =1/δN ≈20s

space scale x = 1mm¯

time scale t = 40s¯

doubling time τ2=ln 2/r ≈ 2h

degradation rate of the chemoattractant a = 5 ⋅ 10−3mol ⋅ s−1 production rate of the chemoattractant b = 4 ⋅ 105cell−1s−1 consumption rate of the nutrient by the bacteria c = 2 ⋅ 10−7cell−1s−1 diffusion coefficient of the nutrient molecules DN =8 ⋅ 10−6cm2⋅s−1 diffusion coefficient of the chemoattractant molecules DS=8 ⋅ 10−6cm2⋅s−1

4. Numerical examples

In this section, we present a large variety of tests in 2dx and one dimensional in velocity space showing the validation of mathematical model and the effectiveness of our numerical schemes. In Test 1, we consider cell aggregation to study only chemotactic motility [18]. In Test 2, we consider wave propagation formed by cells reorientation and the presence of nutrients in a disc. In Test 3, we study the interaction of two traveling wave in a U shape. Then in Test 4, we consider again the traveling wave but in a more wide U shape, which is similar like the effectiveness of centrifugal force. Finally in Test 5 we add the effect of cell division for long time behavior using new parameters.

4.1. Test 1 : cell aggregation. Motile bacteria are able to interact with their environment by accumulating in regions of high concentrations of certain chemicals called attractants and avoiding others called repellents. Cell division is not required to generate these multi-cellular structures because they form on a much shorter time scale than the cell-doubling time [18]. The process leading to formation of multi-cellular clusters can be qualitatively understood as follows: fluctuations in the local cell density produce local gradients of attractants. Cells respond by moving up these concentration gradients thus amplifying the initial spatial non-uniformities in the cell distribution and forming multi-cellular clusters.

We thus use the model of Othmer-Dunbar-Alt [22] to mimic the cell aggregation as follows ∂tf + v ⋅ ∇xf =∫ v′T [S](v, v ′ )f (t, x, v′)dv′− ∫ v′T [S](v ′ , v)f (t, x, v)dv′ , where the tumbling kernel T [S](v′, v) describes the frequency of reorientation v

→v T [S](v′

, v) = ψS(∂tlog(S) + v′⋅ ∇xlog(S)) ,

and the chemical signal is secreted by the cells, following the reaction-diffusion equation (2.7). We use a square domain of size [−0.25, 0.25]2, which is uniformly divided by a mesh size of nx×ny =120 × 120. The velocity space belongs to the unit circle S1, and is uniformly divided into nv =64 parts. All the parameters are chosen as in [28] and listed in Table 1. The initial chemoattractant S0 is equal to 0, and the initial distribution function f0 is independent of the velocity v

f0(x, v) = 100m

(11)

The evolution of the cell aggregation is presented in Figure 3. We observe that the initial density is a Gaussian distribution centered at (0, 0). At time t = 0.2 ¯t, the density forms a volcano profile, which disappears soon and is becoming into an exponential function at time t = 1 ¯t. Finally at time t = 10 ¯t, a steady exponential function is formed. The last observation is similar with the sharp boundary of clusters in [18].

-0.2 -0.1 0 0.1 0.2 -0.2 -0.1 0 0.1 0.2 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 −0.2 −0.1 0 0.1 0.2 0 0.2 0.4 0.6 0.8 1 1.2 −0.2 −0.1 0 0.1 0.2 10−4 10−3 10−2 10−1 100 101 102 t = 0 t = 0 t = 0 -0.2 -0.1 0 0.1 0.2 -0.2 -0.1 0 0.1 0.2 0 0.2 0.4 0.6 0.8 1 1.2 −0.2 −0.1 0 0.1 0.2 0 0.2 0.4 0.6 0.8 1 1.2 −0.2 −0.1 0 0.1 0.2 10−4 10−3 10−2 10−1 100 101 102 t = 0.2 ¯t t = 0.2 ¯t t = 0.2 ¯t -0.2 -0.1 0 0.1 0.2 -0.2 -0.1 0 0.1 0.2 0 5 10 15 20 25 30 35 −0.2 −0.1 0 0.1 0.2 0 5 10 15 20 25 30 35 40 −0.2 −0.1 0 0.1 0.2 10−4 10−3 10−2 10−1 100 101 102 t = 1 ¯t t = 1 ¯t t = 1 ¯t -0.2 -0.1 0 0.1 0.2 -0.2 -0.1 0 0.1 0.2 0 10 20 30 40 50 60 −0.2 −0.1 0 0.1 0.2 0 10 20 30 40 50 60 −0.2 −0.1 0 0.1 0.2 10−4 10−3 10−2 10−1 100 101 102 t = 10 ¯t t = 10 ¯t t = 10 ¯t (a) (b) (c)

Figure 3. Test 1 : Time evolution of the cell aggregation: (a) cell density in domain, (b) section plot of cell density, (c) section plot in log-scale of cell density.

(12)

of cells comprising them. In fact, following [27] the steady density has a form ρ(x) ≃ ρ0exp(−λ∣x∣), when ∣x∣ → ∞,

where ρ0 is the maximum density, λ depends on χS. Notice that the typical size of a cluster, defined as the mean radius given by

< ∣x∣ > = ∫x∣x∣ρ(x)dx ∫xρ(x)dx

= 1 λ

is independent of the total mass. This computation shows that the size of a cluster only depends on the model parameters and is entirely independent of the initial mass condition.

−0.2 −0.1 0 0.1 0.2 0 50 100 150 200 250 Density Radius mass = m mass = 2m mass = 4m −0.2 −0.1 0 0.1 0.2 10−4 10−3 10−2 10−1 100 101 102 103 Density Radius mass = m mass = 2m mass = 4m (a) (b)

Figure 4. Test 1 : The steady-state size of clusters comparison for different total masses at time t = 10 ¯t: (a) section plot of cell density, (b) section plot in log-scale of cell density.

4.2. Test 2 : wave propagation in a disc. Suspended Escherichia coli bacteria swim in convection-free geometries such as capillaries or micro-channels, collectively migrate towards nutrient-rich regions, in the form of propagation concentration waves [1]. In this test, we are interested in the model (2.1)-(2.8) to study concentration waves attracted by nutrient. Moreover, we consider a disc geometry to verify the numerical discretization of boundary conditions in section 3.3.

The computational domain in space is a square domain of size [−3, 3]2, which is uniformly divided by a mesh size of nx×ny = 80 × 80. The disc is inside of the square with radius equal to 3. It is clear that the boundary is not located on grids. Thus to achieve interior high order scheme, some artificial values on ghost points behind the boundary are needed. These values are given by using the numerical method presented in section 3.3. Moreover the velocity space belongs to the unit circle S1, and is uniformly divided into nv=64 parts. All the parameters are chosen as in [28] and listed in Table 1. The initial distribution function f0 is given by a Gaussian function

f0(x, v) = ρ0exp(∣x∣2),

where ρ0 is constant. The initial chemoattractant S0 is equal to 0 and the initial nutrient N0 is homogeneous equal to 1.

(13)

Figure 5). In the third row of Figure 5, the circle wave contracts back to disc center, and finally the cells concentrates at the disc center. We notice that cell diffusion appears when circle wave goes back to disc center. In fact, this diffusion is due to the stiffness of the response functions in the tumbling kernel [28].

-3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 -3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 -3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 t = 0 t = 4 ¯t t = 8 ¯t -3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 -3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 -3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 t = 12 ¯t t = 15 ¯t t = 19 ¯t -3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 -3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 -3 -2 -1 0 1 2 3 x-axis -3 -2 -1 0 1 2 3 y-axis 0 0.5 1 1.5 2 t = 23 ¯t t = 30 ¯t t = 33 ¯t

Figure 5. Test 2 : Time evolution of the cell density at different time.

4.3. Test 3 : interaction of traveling waves in a U shape. In this test, we focus on the influence of the reorientation on the shape. The simulations are compared to a particular experiment by courtesy of Axel Buguin, Institut Curie (see the second row of Figure 9). We consider a channel of U shape, with initially homogeneous nutrient injected in the channel. Two clusters of bacteria are then imposed at two extremities of the channel. These two clusters move along the channel and finally meet at the channel center (the top of U shape). We note that these two clusters keep their bar shape till their meeting.

(14)

divided into nv =64 parts. The numerical parameters used in the simulations are listed in Table 1. The initial distribution function f0 is given by a constant as follows

f0(x, v) = ⎧ ⎪ ⎪ ⎨ ⎪ ⎪ ⎩ 0.25, if 0 ≤ y ≤ 1, 0, else.

The initial chemoattractant S0 is equal to 0 and the initial nutrient N0 is homogeneous equal to 1.

The numerical simulations are presented in the first row of Figure 9. In Figure 9(a), we observe that two bar shape clusters form and are moving to the channel center. In Figure 9(b), two clusters go forward along the half-ring channel and keep well their bar shape. Finally in Figure 9(c), two clusters meet at the channel center. We see that the numerical simulations have a good agreement with the experiment.

0 1 2 3 4 5 6 7 8 x-axis 0 1 2 3 4 5 6 y-axis 0 0.5 1 1.5 2 2.5 3 3.5 0 1 2 3 4 5 6 7 8 x-axis 0 1 2 3 4 5 6 y-axis 0 0.5 1 1.5 2 2.5 3 3.5 0 1 2 3 4 5 6 7 8 x-axis 0 1 2 3 4 5 6 y-axis 0 0.5 1 1.5 2 2.5 3 3.5 (a) (b) (c)

Figure 6. Test 3 : Time evolution of the cell density (a) t= 0.65 sec. (b) t = 2.65 sec. (c) t = 4.53 sec. At the top numerical results and at the bottom experiments on Escherichia coli (courtesy of Axel Buguin, Institut Curie).

4.4. Test 4 : one traveling wave in a U shape. In the last test, we consider again traveling wave in a U shape but with a more wide channel than the previous one. The experience is shown in the second row of Figure 7 and Figure 8. Again we inject homogeneous nutrient in the channel, but we consider only one cluster of bacteria at the right extremity of the channel. We observe that at straight part of channel the cluster goes ahead in a bar shape. Once the cluster enters into the half-ring part, the bacteria near interior circle goes faster than the one near the exterior circle. Moreover, bacteria contracts towards the exterior circle. Before the cluster enters into the left straight part of channel, bacteria concentrate almost near the exterior circle. When the cluster goes forward the other extremity, the cluster recovers its original bar shape.

(15)

The numerical simulations are presented in the first row of Figure 7 and Figure 8. We can see that the numerical simulations are very similar as the experience one. In fact, this phenomena is due to the directional persistence of chemotactic bacteria in a traveling con-centration wave. When bacteria enter into a wide half-ring, they keep going straight and accumulate by the reflection of the exterior circle. It is very similar like the effectiveness of centrifugal force, and can be observed significantly in a wide channel. This test shows that the model (2.1)-(2.8) represents well the chemotactic bacteria behavior and our numerical discretization based on Cartesian mesh approximates well the continuous model.

(a) (b) (c)

Figure 7. Test 4 : Numerical simulations (top) and experiments on Es-cherichia coli (bottom) : time evolution of the cell density (a) t = 7.5 t (b) t = 18.5 t (c) t = 23 t in sec.

4.5. Test 5 : long time behavior and pattern formations. In this last numerical test, we consider the full kinetic model with cell divisions and degradation

(4.1) ∂f

∂t +v ⋅ ∇xf = Q(f ) + rf − γΘ(ρ) Θ(ρ − ρ∞), x ∈ Ω, v ∈ V, where the division rate is given by [26]

r = G0N σ + N,

(16)

(d) (e) (f) Figure 8. Test 4 : Numerical simulations (top) and experiments on Es-cherichia coli (bottom) : time evolution of the cell density (d) t = 28 t (e) t = 34.5 t (f ) t = 46.5 t in sec.

Actually such a source term has been proposed in [26] in the framework of a macroscopic Patlak-Keller-Segel model where the unknown is the total density ρ. Numerical simulations of this macroscopic model have shown pattern formations as the ones observed in experiments [26,6]. Here we consider the following initial density

f0(x, v) ∶= { 1 if ∣x∣ ≤ 1, 0 else

and the initial nutrient concentration is uniform N0=0.5. In the nutrient and chemoattrac-tant system of equations (2.7) we take the following parameters DS=DN =1, the production rate of chemoattractant b = 20, the degradation rate of chemoattractant a = 8 and the con-sumption rate of nutrient c = 0.8, whereas in (4.1), we choose σ = 0.1, ρ∞ =15 and for the turning operator (2.2)-(2.4),we have δ = 20, ψ0 =1, χN =1/2 and χS=1/10. The numerical simulations are presented in Fig. 9. We observe that for high initial nutrient concentration, cell density in the expanding ring becomes sufficient both for the break of stability of the uniform cell distribution and for their aggregation. In particular, if after the formation of the successive set of aggregates the expanding ring had time to grasp a certain part of the cells, participating in aggregation, then a radial pattern is formed.

5. Conclusion

(17)

Figure 9. Test 5 : Time evolution of the cell density (a) t = 0 t (b) t = 15 t (c) t = 25 t (d) t = 35 t (e) t = 45 t and (f ) t = 55 t in sec.

chemical substances, i.e. nutrient and chemottractant. Then we give the numerical discretiza-tion for this kinetic model and the numerical method for the boundary condidiscretiza-tions based on a Cartesian mesh. A large various numerical tests in 2Dx×1Dv are shown and compared with biological experiences. We conclude that on the one hand this kinetic model represents well the chemotactic bacteria behaviors and on the other hand our numerical method is accurate and efficient for numerical simulation.

References

[1] J. Adler, Chemotaxis in bacteria, Science 153:708-716

[2] W. Alt, Biased random walk models for chemotaxis and related diffusion approximations, J. Math Biol. 9 pp. 147-177

[3] E. Bouin, V. Calvez and G. Nadin, Hyperbolic traveling waves driven by growth, preprint (2011). [4] S.M. Block, J.E. Segall and H.C. Berg, Adaptation kinetics in bacterial chemotaxis, J. Bacterial 154

(1983) pp. 312-323

[5] N. Bournaveas and V. Calvez, A review of recent existence and blow-up results for kinetic models of chemotaxis, Can. Appl. Math. Q. (2010).

[6] E. Budrene and H. Berg, Complex patterns formed by motile cells of Escherichia coli,Nature, 349:630-633. [7] C. Cercignani, The Boltzmann equation and its applications, Springer-Verlag, Berlin (1988)

[8] A. Chertock and A. Kurganov, A second-order positivity preserving central-upwind scheme for chemotaxis and haptotaxis models. Numer. Math. 111 (2008), no. 2, 169205.

[9] Y. Epshteyn and A. Izmirlioglu, Fully discrete analysis of a discontinuous finite element method for the Keller-Segel chemotaxis model. J. Sci. Comput. 40 (2009), pp. 211256

[10] R. Erban and H.G. Othmer, From individual to collective behavior in bacterial chemotaxis, SIAM J. Appl. Math. 65 (2004), pp. 361391.

(18)

[12] F. Filbet and C.-W. Shu Approximation of hyperbolic models for chemosensitive movement, SIAM J. Sci. Comput.

[13] F. Filbet, A finite volume scheme for the Patlak-Keller-Segel chemotaxis model. Numer. Math. 104 (2006), pp. 457488.

[14] F. Filbet and C. Yang, Inverse Lax-Wendroff method for boundary conditions of Boltzmann equations, Submitted

[15] Y. V. Kalinin, L. Jiang, Y. Tu and M. Wu, Logarithmic sensing in Escherichia coli bacterial chemotaxis, Biophys J., 96 pp. 2439-2448, (2009)

[16] R. M. Macnab and D. E. Koshland Jr, The gradient-sensing mechanism in bacterial chemotaxis, Proc. Natl. Acad. Sci. USA 69 (1972), 25092512.

[17] A. Marrocco; 2D simulation of chemotaxis bacteria aggregation, ESAIM:M2AN 37, No.4, (2003) pp. 617–630

[18] Nikhil Mittal, Elena O. Budrene, Michael P. Brenner, and Alexander van Oudenaarden, Motility of Escherichia coli cells in clusters formed by chemotactic aggregation, Proceedings of the National Academy of Sciences of the United States of America, 100 (2003)

[19] G.-S. Jiang and C.-W. Shu, Efficient implementation of weighted ENO schemes, Journal of computa-tional physics 126, 202–228 (1996).

[20] E. F. Keller, L.A. Segel, Traveling bands of chemotactic bacteria: a theoretical analysis, J. Theor. Biol. 30 pp. 235-248, (1971)

[21] R. Natalini and M. Ribot, Asymptotic high order mass-preserving schemes for a hyperbolic model of chemotaxis. SIAM J. Numer. Anal. 50 (2012), pp. 883-905.

[22] H. G. Othmer, S. R. Dunbar, W. Alt, Models of dispersal in biological systems., J Math Biol. 26 pp. 263-98. (1988)

[23] C.S. Patlak, Random walk with persistense and external bias. Bull. Math. Biol. Biophys. 15, (1953) pp. 311–338

[24] B. Perthame, PDE models for chemotactic movements: parabolic, hyperbolic and kinetic, Appl. Math. 49 (2004), 539564.

[25] B. Perthame, Transport Equations in Biology, Frontiers in Mathematics, Birkh¨auser Verlag, Basel, 2007. [26] A.A. Polezhaev, R.A. Pashkov, A.I. Lobanov, I.B. Petrov, Spatial patterns formed by chemotactic bacteria

Escherichia coli Int. J. Developmental Biology, 50, pp. 309-314 (2006)

[27] J. Saragosti, V. Calvez, N. Bournaveas, A. Buguin, P. Siberzan and B. Perthame, Mathematical Description of Bacterial Traveling Pulses, PLoS Comput. Biol. 6:e1000890

[28] J. Saragosti, V. Calvez, N. Bournaveas, B. Perthame, A. Buguin and P. Siberzan, Directional persistence of chemotactic bacteria in a traveling concentration wave, Proceedings of the National Academy of Sciences of the United States of America, (2011)

[29] P.A. Spiro, J.S. Parkinson and H.G. Othmer, A model of excitation and adaptation in bacterial chemo-taxis, Proc. Natl. Acad. Sci. USA 94 (1997), 72638.

[30] S. Tan and C.-W. Shu, Inverse Lax-Wendroff procedure for numerical boundary conditions of conservation laws, Journal of Computational Physics, 229 (2010), 8144–8166.

[31] B. van Leer, Towards the ultimate conservative difference scheme II. Monotonicity and conservation combined in a second order scheme, J. Comput. Phys., 14 (1974), pp. 361370.

Références

Documents relatifs

We also prove the following statement, Theorem 2, which may be interpreted as follows: the best order of convergence, for the weak error, when considering the supremum over all

Here, the contribution of anisotropic unsteady mesh adaptation to this issue is discussed, which aims at increasing the accuracy of the solution while decreasing the CPU time of

Unstructured mesh, continuous mesh, Riemannian metric space, interpolation error, linear interpolate, anisotropic mesh adaptation, optimal interpolation error bound.. AMS

We have presented in this section two numerical schemes for an hyperbolic semi- linear system of chemotaxis (5) : an asymptotic high order scheme and a well-balanced scheme.. The

This opens the possibility of explaining veridical OBE reports in terms of the fact that the individual left her body and observed the world from a position external to

In Section 4.1, the possibilities of a simple continuous bond of both sides of the switching con- straint is formalised by the switching constraint connected (SCC) relaxation;

Thacker test case: Comparison of orders of error for the water height h (left) and the discharge q (right) for different schemes at time T =

Then, in the second part of the paper, we describe the kinetic system derived as the limit from the particle model as N tends to infinity; we propose, in 1D, a numerical scheme for